UCLA UCLA Previously Published Works - eScholarship

Page created by Martin Lowe
 
CONTINUE READING
UCLA UCLA Previously Published Works - eScholarship
UCLA
UCLA Previously Published Works

Title
Computational Exploration of Ambiphilic Reactivity of Azides and Sustmann's Paradigmatic
Parabola.

Permalink
https://escholarship.org/uc/item/7wg7936n

Journal
The Journal of organic chemistry, 86(8)

ISSN
0022-3263

Authors
Chen, Pan-Pan
Ma, Pengchen
He, Xue
et al.

Publication Date
2021-04-01

DOI
10.1021/acs.joc.1c00239

Peer reviewed

 eScholarship.org                               Powered by the California Digital Library
                                                                University of California
UCLA UCLA Previously Published Works - eScholarship
pubs.acs.org/joc                                                                                                                                    Article

Computational Exploration of Ambiphilic Reactivity of Azides and
Sustmann’s Paradigmatic Parabola
Pan-Pan Chen,§ Pengchen Ma,§ Xue He, Dennis Svatunek, Fang Liu,* and Kendall N. Houk*
      Cite This: J. Org. Chem. 2021, 86, 5792−5804                                     Read Online

ACCESS                        Metrics & More                              Article Recommendations                 *
                                                                                                                  sı Supporting Information

ABSTRACT: We examine the theoretical underpinnings of the
seminal discoveries by Reiner Sustmann about the ambiphilic
nature of Huisgen’s phenyl azide cycloadditions. Density functional
calculations with ωB97X-D and B2PLYP-D3 reproduce the
experimental data and provide insights into ambiphilic control of
reactivity. Distortion/interaction-activation strain and energy
decomposition analyses show why Sustmann’s use of dipolarophile
ionization potential is such a powerful predictor of reactivity. We
add to Sustmann’s data set several modern distortion-accelerated
dipolarophiles used in bioorthogonal chemistry to show how these
fit into the orbital energy criteria that are often used to understand
cycloaddition reactivity. We show why such a simple indicator of
reactivity is a powerful predictor of reaction rates that are actually
controlled by a combination of distortion energies, charge transfer, closed-shell repulsion, polarization, and electrostatic effects.

■    INTRODUCTION
Sustmann and Trill published a landmark paper in 1972 about
                                                                                    The Sustmann paper quantified what we now describe as the
                                                                                 ambiphilic character of aryl and alkyl azides, a feature that was
reactivity of phenyl azide in 1,3-dipolar cycloadditions.1a At that              revealed by comparisons of Huisgen’s experimental measure-
time, an iconic parabola (Figure 1a) presented for the first time a               ments3 to the many ionization potentials that became available
memorable summary of how the electronic nature of a                              from Heilbronner et al.4 and Sustmann et al.’s work,5 as well as
dipolarophile, relative to that of the 1,3-dipole, determines the                ours6 and others7 through the commercialization of the
rate of reaction. This parabola, based on second-order                           PerkinElmer UV photoelectron spectrometer in 1967.8
perturbation theory,2 shows the correlation of the second-                          Ever since that time, Fukui’s FMO theory has provided a
order rate constant k for cycloadditions of phenyl azide to a                    powerful model for a qualitative understanding of reactivity and
series of dipolarophiles and the experimental ionization                         regioselectivity in reactions of various 1,3-dipoles with alkenes.9
potential (IP) of these dipolarophiles.1b The IP is related by                   We recently proposed a general distortion/interaction (D/I)
Koopmans’ theorem to the negative of the highest occupied                        model for cycloaddition reactivity,10a−d and Bickelhaupt
molecular orbital (HOMO) energy and serves as a measure of                       proposed the equivalent activation strain (AS) model.10e,f The
both the HOMO and lowest unoccupied molecular orbital                            D/I-AS model has been applied to 1,3-dipolar cycloadditions
(LUMO) energies: a low-energy HOMO is usually accom-                             and shows that there are correlations between distortion
panied with a low-energy LUMO and vice versa.                                    energies and the activation barriers for 1,3-dipolar cyclo-
   Figure 1b shows how donor (D) and acceptor (A)                                additions of various 1,3-dipoles with alkenes and alkynes. The
substituents alter the frontier molecular orbital (FMO) energies                 reactivity differences between dipoles are often controlled by the
of a dipolarophile.1c In the reaction of phenyl azide with ethylene              distortion energies of the 1,3-dipoles (the energy required to
(Figure 1b, middle), two sets of interactions contribute: the                    distort the ground state of the 1,3-dipoles to their transition state
HOMO of azide with the LUMO of ethylene and the HOMO of                          geometries), rather than by FMO interactions or reaction
ethylene with the LUMO of azide. An electron donor (Figure                       thermodynamics. However, the reactivity of dipolarophiles with
1b, left) raises the ethylene HOMO and LUMO energies, which
leads to better interaction between the HOMO of ethylene and
the LUMO of azide, resulting in stronger stabilization and higher                Received: January 30, 2021
reactivity. An electron acceptor (Figure 1b, right) enhances the                 Published: March 26, 2021
reactivity of ethylene in the same manner by lowering the
HOMO and LUMO energies. With this FMO model, Sustmann
qualitatively explained the effect of different substituents on
reactivities in 1,3-dipolar cycloadditions.
                                       © 2021 The Authors. Published by
                                             American Chemical Society                                                   https://doi.org/10.1021/acs.joc.1c00239
                                                                          5792                                                J. Org. Chem. 2021, 86, 5792−5804
UCLA UCLA Previously Published Works - eScholarship
The Journal of Organic Chemistry                                                   pubs.acs.org/joc                                                 Article

Figure 1. (a) Plot of reactivity (lnk2) vs IP. Replotted in the same style as Sustmann’s original figure in ref 1. Copyright 1972 by Verlag Chemie GmbH,
Germany. (b) Schematic of the azide HOMO and LUMO vs those of donor-substituted dipolarophiles, ethylene, and acceptor-substituted
dipolarophiles.

                                                                         5793                                            https://doi.org/10.1021/acs.joc.1c00239
                                                                                                                              J. Org. Chem. 2021, 86, 5792−5804
UCLA UCLA Previously Published Works - eScholarship
The Journal of Organic Chemistry                                                 pubs.acs.org/joc                                                 Article

Scheme 1. The 1,3-Dipolar Cycloaddition between Phenyl Azide and Different Dipolarophiles Studied by Sustmann and in This
Paper

Table 1. Experimental Rate Constantsa and Calculated Reaction Barriers and Rate Constantsb
                                                                   method Ac                                          method Bd
    dipolarophile           log(kexp × 109)         ΔGA‡ (kcal/mol)            log(kcal × 109)         ΔGB‡ (kcal/mol)               log(kcal × 109)
           e
        R1                       7.06                     23.1                      4.85                     18.3                          8.37
        R2e                      6.00                     25.6                      3.02                     20.8                          6.54
        R3e                      5.41                     24.3                      3.97                     19.6                          7.42
        R4e                      4.40                     29.6                      0.09                     23.7                          4.41
        R5                       4.29                     30.0                     −0.21                     23.9                          4.27
        R6e                      2.38                     32.5                     −2.04                     26.4                          2.43
        R7                       1.95                     32.9                     −2.33                     26.7                          2.21
        R8                       1.60                     32.3                     −1.90                     27.8                          1.41
        R9                       1.38                     33.8                     −3.00                     28.4                          0.97
        R10                      1.18                     33.3                     −2.63                     27.6                          1.55
        R11                      0.52                     35.0                     −3.88                     28.9                          0.60
        R12                      1.43                     33.4                     −2.70                     27.9                          1.33
        R13                      1.46                     33.5                     −2.78                     27.7                          1.48
        R14                      1.53                     32.5                     −2.04                     26.3                          2.51
        R15                      1.60                     33.3                     −2.63                     27.0                          1.99
        R16                      1.86                     33.4                     −2.70                     27.3                          1.77
        R17                      2.03                     32.7                     −2.19                     28.0                          1.26
        R18                      2.92                     31.7                     −1.46                     25.9                          2.80
        R19                      2.99                     31.5                     −1.31                     26.4                          2.43
        R20                      3.02                     31.5                     −1.31                     26.6                          2.29
        R21                      3.40                     31.3                     −1.16                     25.6                          3.02
a
 For different dipolarophiles, the corresponding cycloaddition was performed in a particular solvent (carbon tetrachloride or benzene) at 25 °C.30
b
 For different dipolarophiles, the calculated reaction barriers and rate constants are values in a specific solvent (carbon tetrachloride or benzene),
which was chosen according to what was used in the experiment.30 cωB97X-D/aug-cc-pVTZ-CPCM(solvent)//ωB97X-D/6-31+G(d,p)-
CPCM(solvent). dB2PLYP-D3/aug-cc-pVTZ-CPCM(solvent)//ωB97X-D/6-31+G(d,p)-CPCM(solvent). eIn the experimental study, the
reaction solvent was benzene for these dipolarophiles and carbon tetrachloride for the other dipolarophiles.30

a given dipole correlates with orbital interactions and can be                any discussion of reactivity in terms of orbitals.11 The MEDT
understood with FMO analysis.10a Energy decomposition                         approach was applied to azide cycloadditions and many other
analyses have shown that a variety of other factors, like closed-             cycloadditions.11b In addition, Cremer et al. performed
shell repulsion, electrostatic effects, and polarization, may also             computational studies of ten different cycloadditions of 1,3-
influence reactivity.10e A different approach based on electron                 dipoles XYZ (diazonium betaines, nitrilium betaines, azome-
density analyses was proposed by Domingo.11 Domingo’s                         thines, and nitryl hydride) to acetylene with B3LYP/6-31G(d,p)
approach, molecular electron density theory (MEDT), is                        geometries and CCSD(T)-F12/aug-cc-pVTZ energetics, em-
based on conceptual density functional theory, in which                       ploying their unified reaction valley approach (URVA)12 on the
reactivity parameters are obtained by computations of changes                 concerted processes, and Breugst, with Huisgen and Reissig,
in electron density upon electron removal or addition. This                   conducted calculations with M06-2X geometries and DLPNO-
obviously ties in closely with Sustmann’s ionization potential as             CCSD(T)13 energies for the concerted 1,3-dipolar cyclo-
an indicator of reactivity, although Domingo criticizes severely              additions of diazomethane with many model-substituted
                                                                       5794                                           https://doi.org/10.1021/acs.joc.1c00239
                                                                                                                           J. Org. Chem. 2021, 86, 5792−5804
UCLA UCLA Previously Published Works - eScholarship
The Journal of Organic Chemistry                                                      pubs.acs.org/joc                                                  Article

alkynes.14 Because few recent high-accuracy calculations have
explored the full range of azide cycloadditions to the alkene
dipolarophiles studied experimentally by Huisgen, we carried
out an analysis of the Sustmann relationship for this important
class of reactions.15

■   COMPUTATIONAL METHODS
All density functional theory (DFT) calculations were performed with
Gaussian 16.16 Geometry optimizations of all stationary points were
performed with the long-range corrected hybrid functional, ωB97X-
D17 with the 6-31+G(d,p) basis set in solution. Solvents were chosen
according to what was used in the experiment. Vibrational frequencies
were calculated at the same level of theory to evaluate the zero-point
vibrational energy (ZPVE) and thermal corrections at 298.15 K. The
single-point energies were computed using ωB97X-D and the double
hybrid B2PLYP-D318 functional with the aug-cc-pVTZ19 basis set;
solvation energy corrections were evaluated with the CPCM model.20
The Hirshfeld charges of the transition states were analyzed at the
ωB97X-D/6-31+G(d,p) level of theory on the optimized structures to
determine the direction and extent of charge transfer. Frontier
molecular orbitals were calculated on DFT-optimized structures at                 Figure 3. Correlation of logarithms of the experimental rate constants
the HF level of theory with the 6-31G(d) basis set and visualized with            with the calculated rate constants (method B) for the cycloadditions
Multiwfn21 and VMD.22 Fragment distortion and interaction energies                between phenyl azide and 21 dipolarophiles.
were calculated with autoDIAS23 at the ωB97X-D level of theory with
the aug-cc-pVTZ basis set in the gas phase. Energy decomposition
analyses were performed in ADF 2019.30424 at the ωB97X-D3/
TZ2P25 level of theory using PyFrag 2019.26 The orbital coefficients
were calculated using the NBO27 module of Gaussian 16 and Multiwfn.
Extensive conformational searches for the intermediates and transition
states have been carried out to ensure that the lowest energy conformers
were located. The 3D images of molecules were generated using
CYLView.28

■    RESULTS AND DISCUSSION
We investigated the reactivity of phenyl azide in the context of
modern density functional theory and the distortion/inter-
action-activation strain model and analyzed the interactions of
phenyl azide and dipolarophiles with an energy decomposition
analysis. We describe the factors that control reactivity and show
why Sustmann’s parabola provides such a powerful model for

                                                                                  Figure 4. Theoretical version of Sustmann’s plot with charge transfer
                                                                                  (e), using B2PLYP-D3/aug-cc-pVTZ-CPCM(solvent)//ωB97X-D/6-
                                                                                  31+G(d,p)-CPCM(solvent)-predicted rate constants and HF/6-
                                                                                  31G(d)-calculated ionization potentials (eV). The three purple points,
                                                                                  for which experimental rate constants have been given in Huisgen’s
                                                                                  experimental study30 but were not included in Sustmann’s analysis, are
                                                                                  qualitatively in line with the parabolic trend. The three points in red are
                                                                                  added to the original plot to show how distortion-accelerated
                                                                                  dipolarophiles fit on the correlation, and the point in orange is added
                                                                                  to the original plot to show the enormous acceleration that occurs with
                                                                                  Cu catalysis. Data in blue (negative number) shows charge transfer
Figure 2. Correlation of aΔGA‡ with bΔGB‡ for the cycloadditions                  from dipolarophiles to azide; data in green (positive number) shows
between phenyl azide and 21 dipolarophiles. aωB97X-D/aug-cc-pVTZ-                 charge transfer from azide to dipolarophiles; data in black shows charge
CPCM(solvent)//ωB97X-D/6-31+G(d,p)-CPCM(solvent);                                 transfer of 0.05 e or less between reactants, essentially zero.
b
  B2PLYP-D3/aug-cc-pVTZ-CPCM(solvent)//ωB97X-D/6-31+G-
(d,p)-CPCM(solvent).

                                                                           5795                                              https://doi.org/10.1021/acs.joc.1c00239
                                                                                                                                  J. Org. Chem. 2021, 86, 5792−5804
UCLA UCLA Previously Published Works - eScholarship
The Journal of Organic Chemistry                                                     pubs.acs.org/joc                                             Article

Figure 5. Plots of logarithms of the experimental rate constants versus (a) distortion and (b) interaction energies for the cycloadditions between
phenyl azide and dipolarophiles. Energies are in kcal/mol.

                                                                                 reflect corresponding shifts of LUMO energy. We indeed found
                                                                                 a correlation between predicated ionization potential and
                                                                                 electron affinity (EA) for different dipolarophiles (Figure S2),
                                                                                 indicating that the change of LUMO energy can be semi-
                                                                                 quantitatively characterized by the variation of HOMO energy.
                                                                                    We also plotted the predicted log k values for the 1,3-dipolar
                                                                                 cycloadditions versus the calculated ionization potentials of a
                                                                                 number of dipolarophiles. Figure 4 is a theoretical version of
                                                                                 Sustmann’s plot. The parabolic relationship of reactivity with
                                                                                 dipolarophile ionization potential is nicely consistent with
                                                                                 Sustmann’s results.1a Additionally, we plotted the relationship
                                                                                 between logarithms of the predicted rate constants of 1,3-dipolar
                                                                                 cycloadditions and calculated Mulliken electronegativities (χ)
                                                                                 for all dipolarophiles, and we obtained a similar parabola (Figure
                                                                                 S3). This indicates that there is a relationship between the
                                                                                 ionization potential used by Sustmann and the electronegativity,
                                                                                 which is the average of the ionization energy and the electron
                                                                                 affinity (Figure S4). It is worth noting that the correlation with
Figure 6. Plot of optimum FMO gaps (eV) between dipole and                       Mulliken’s electronegativity is only a slight modification of the
dipolarophiles versus calculated ionization potentials (eV) of a number          poor correlation between IP and EA (compare Figures S2 and
of dipolarophiles.                                                               S4) because Mulliken’s electronegativties (χ = IP/2 + EA/2)
                                                                                 superimpose half of the EA values on the identity correlation IP/
ambiphilic cycloaddition reactivity, despite the many factors that               2 vs IP and thus reduce the magnitude of the deviations while
contribute to reactivity.                                                        simultaneously increasing the slope.
    We computed the transition states of the 1,3-dipolar                            We have, in addition to the 21 dipolarophiles reported in
cycloadditions between phenyl azide and 27 dipolarophiles.                       Sustmann’s analysis, studied the cycloadditions of phenyl azide
The first 21 dipolarophiles, which are included in Sustmann’s                     with other dipolarophiles, such as 1-methoxycyclopentene,32
plot, are listed in Scheme 1. Table 1 shows the energetics                       maleic anhydride, and N-phenylmaleimide, for which exper-
computed with two functionals (method A and method B). The                       imental rate constants were given in Huisgen et al.’s 1967
same trend of reactivity was obtained from the two methods                       paper.30 The reactivities for these three species (Figure 4,
(Figure 2). The computed rate constants (by applying the                         purple) are qualitatively in line with the parabolic trend.
Eyring equation to activation free energies)29 given by the                      Moreover, we have incorporated the results for distortion-
double hybrid functional (method B) agree somewhat better                        activated (or strain-promoted) dipolarophiles,33 e.g., trans-
with measured rate constants30 (Table 1 and Figure 3). For all                   cyclooctene, cyclooctyne, and cyclopropene, which are of
dipolarophiles studied, there is a good linear correlation between               interest in bioorthogonal chemistry (Figure 4, red).34 It is
the variables. In the following discussion, we report the values of              shown that trans-cyclooctene actually fits very nicely on the
energetics calculated by B2PLYP-D3, unless otherwise specified.                   parabola since the distortion of the alkene increases the HOMO
    We compared our predicted ionization potentials (IP) based                   energy to be about the same as pyrrolidinocyclohexene (R2).
on Koopmans’ theorem31 to those measured experimentally by                       However, cyclooctyne has a slightly lowered ionization potential
photoelectron spectroscopy and found that the theoretical value                  (9.4 eV) as compared to 2-butyne (9.6 eV)35 but is anomalously
is in good agreement with the experimental value (Figure S1). In                 reactive, as expected for a distortion-accelerated dipolaro-
Sustmann’s plot, a key concept is that the HOMO and LUMO of                      phile.33,36 Cyclopropenes have been used as dienophiles in
the substituted dipolarophiles will be shifted in the same                       bioorthogonal chemistry by the Prescher34d and Devaraj
direction. Therefore, the variation of HOMO energy should                        groups34e and studied theoretically.34f Cyclopropene is some-
                                                                          5796                                         https://doi.org/10.1021/acs.joc.1c00239
                                                                                                                            J. Org. Chem. 2021, 86, 5792−5804
UCLA UCLA Previously Published Works - eScholarship
The Journal of Organic Chemistry                                                 pubs.acs.org/joc                                              Article

Figure 7. DFT-optimized transition structures for cycloadditions between phenyl azide and five representative dipolarophiles. TS1 and TS17 are for
the experimentally observed regioselectivity.

what more reactive than reactive norbornene. To support our                  dipolarophiles, the HOMO-LUMO interaction is weak (larger
statement that trans-cyclooctene, cyclooctyne, and cyclo-                    FMO gap), so the transition state is later, corresponding to
propene are all distortion-promoted dipolarophiles, we                       greater distortion. In other words, the orbital interactions are
performed distortion/interaction-activation strain analysis on               related to the transition state position and therefore the
the transition states of their cycloaddition reactions with phenyl           distortion energies. Although Sustmann’s parabola provided a
azide (Figure S5). The corresponding results confirmed our                    great qualitative guide to the reactivity trends for an ambiphilic
inference (Table S1). Finally, we add the famous Sharpless Cu-               1,3-dipole, it is because the IP reflects not only changes in
catalyzed click reaction of azide with terminal alkyne (Figure 4,            HOMO or LUMO energy but also distortion energy differ-
orange) to the graph to emphasize the enormous acceleration                  ences.38
that occurs with Cu catalysis.37                                                Figure 6 shows a plot of the optimum FMO gaps between
   We have also calculated the charge transfer for all cyclo-                dipole and 21 dipolarophiles (the FMOs used here are those
addition transition states, and values are shown in Figure 4. In             involved in the formation of new bonds, not the actual HOMOs
general, the fast reactions involve significant (δq > 0.05 e) charge          and LUMOs) versus the calculated ionization potentials of these
transfer, which is consistent with the general idea that orbital             dipolarophiles. There is an excellent linear correlation between
interactions are influential on reactivity.                                   IP and the energy gap. For the electron-rich (R1−R5, red dots)
   We performed distortion/interaction-activation strain anal-               and neutral or ambiphilic dipolarophiles (R6−R13, R15, and
yses on these cycloaddition transition states to evaluate other              R16, green dots), lower IP and a smaller FMO gap correlate with
factors that influence reactivities. This involves decomposition              higher reactivity. However, with the electron-deficient dipolar-
of the electronic energy (ΔE) into two terms: the distortion                 ophiles, R14 and R17−R21 (blue dots), there must be other
energy (ΔEdist) that results from the distortion of the individual           factors that compensate the unfavorable FMO gaps, leading to
reactants and the interaction (ΔEint) between the deformed                   high reactivity.
reactants. Based on our study, we find that there is a rough                     A plot of the IP of the dipolarophiles versus both FMO gaps
relationship between distortion energy and cycloaddition                     (the FMOs used here are those involved in the formation of new
reactivity (Figure 5a). However, the interaction energy                      bonds, not the actual HOMOs and LUMOs) is provided in
surprisingly has little correlation with the reactivity (Figure              Figure S6. As expected, the HOMOdipolarophile-(LUMO+2)azide
5b). This is, of course, in apparent contrast to Sustmann’s                  gap of electron-rich dipolarophiles (IP < 10 eV) accounts for the
conclusion and that from charge transfer assessments that                    left arm of the parabola. For dipolarophiles with IP around 10
suggest that FMO interactions control cycloaddition reactivity.              eV, HOMOdipolarophile-(LUMO+2)azide and (HOMO-3)azide-
As shown in Figure 5a, nucleophilic dipolarophiles (R1−R5, red               LUMOdipolarophile gaps become comparable; for dipolarphiles
dots) show low distortion energies and higher reactivities;                  with IP greater than 11 eV, the (HOMO-3)azide-LUMOdipolarophile
electrophilic dipolarophiles (R14 and R17−R21, blue dots)                    gap becomes determinant and corresponds with the right arm of
show moderate distortion energies and reactivities; neutral or               the parabola. This is the basic idea that Sustmann proposed and
ambiphilic dipolarophiles (R6−R13, R15, and R16, green dots)                 is now quantitatively shown in Figure S6.
show the greatest distortion energies and the lowest reactivities.              To gain more insights into the physical factors leading to the
The cycloaddition reactivity as a whole is related to the                    reactivity trend, we studied the cycloadditions between phenyl
distortion energy. There is a close relationship between FMO                 azide and five representative dipolarophiles, including the most
interactions, the position of the transition state, and distortion           reactive and electron-rich R1 and R4, neutral R11, and electron-
energies. For nucleophilic dipolarophiles, the HOMO-LUMO                     deficient R17 and R21. The corresponding transition states are
interaction is stronger (smaller FMO gap), so the transition state           shown in Figure 7. The forming C−N bond ranges from 2.0 to
is earlier, corresponding to less distortion. For electrophilic              2.8 Å, and the transition state structure varies from nearly
                                                                      5797                                          https://doi.org/10.1021/acs.joc.1c00239
                                                                                                                         J. Org. Chem. 2021, 86, 5792−5804
UCLA UCLA Previously Published Works - eScholarship
The Journal of Organic Chemistry                                                 pubs.acs.org/joc                                                Article

Figure 8. Orbital interaction diagrams for the cycloadditions between phenyl azide and selected dipolarophiles. In order to justify that HOMO-3 and
LUMO+2 of phenyl azide are the effective orbitals involved in the orbital interactions between phenyl azide and dipolarophiles, we show the HOMO,
HOMO-1, and HOMO-2 as well as the LUMO and LUMO+1 of phenyl azide in the upper left corner for comparison.

synchronous to highly asynchronous, with an average of the two                are involved in the formation of new bonds, not the actual
forming C−N bonds being 2.2−2.4 Å.                                            HOMOs and LUMOs. For the cycloaddition between phenyl
  Figure 8 shows the orbital interaction diagrams of the selected             azide and R1, the dominant orbital interaction is between the
reactions. As noted above, these diagrams show the FMOs that                  HOMO of R1 and the LUMO+2 of phenyl azide (the LUMO+2
                                                                       5798                                           https://doi.org/10.1021/acs.joc.1c00239
                                                                                                                           J. Org. Chem. 2021, 86, 5792−5804
UCLA UCLA Previously Published Works - eScholarship
The Journal of Organic Chemistry                                                   pubs.acs.org/joc                                                Article

Figure 9. Distortion/interaction-activation strain (DIAS) analysis for selected cycloaddition reactions. Reactant distortion and interaction energies
were calculated at the ωB97X-D level of theory with the aug-cc-pVTZ basis set, in the gas phase. (a) Electronic energies (ΔE). (b) Distortion energies
(ΔEdist). (c) Interaction energies (ΔEint).

is slightly higher than the perpendicular LUMO, which                          potential and kinetic energies occur as a result. We caution
conjugates with the phenyl group and is not involved in bond                   against simple interpretations of the very large changes in all
formation). This is the same with R4, R11, and R17. Along the                  energy components that occur in partial covalent bond
series, as the HOMO energy of a dipolarophile decreases (the                   formation; nevertheless, details of the energy decomposition
nucleophilicity of the dipolarophile decreases), the energy gap                analysis are given in the Supporting Information (Figure S7).
increases (11.1, 12.7, 12.8, and 14.4 eV for R1, R4, R11, and                     We also investigated the regioselectivity of the 1,3-dipolar
R17, respectively). For R21, the interaction between HOMO-3                    cycloadditions for five unsymmetrical alkenes, R2, R8, R9, R15,
of phenyl azide (the higher energy orbitals are not involved in                and R17. Orbital interaction diagrams for the corresponding
bond formation) and the LUMO of R21 dominates. Phenyl                          cycloaddition reactions are shown in Figure 10. The
azide is somewhat electron-deficient and thus gives stronger                    regioselectivity has been rationalized traditionally by frontier
interactions and faster rates with the electron-rich dipolar-                  molecular orbital interactions, and in general, the preferred
ophiles.                                                                       product is indeed the one in which the favored product results
   A distortion/interaction-activation strain analysis was per-                from the union of the larger terminal coefficients of azide (the
formed along the reaction coordinate (defined by the bond                       nucleophilic substituted terminus of azide in the HOMO-3 and
length of the shorter forming C−N bond) for the five reactions                  the electrophilic unsubstituted terminus in the LUMO+2, see
in discussion (Figure 9). The transition states of cycloadditions              Figure 8) with the larger terminal coefficients in the
between phenyl azide and symmetric dipolarophiles (R4, R11,                    complementary FMO of the dipolarophile.9
and R21) are nearly synchronous, corresponding to generally                       To achieve a more quantitative view, we performed a
higher distortion energies. Asymmetric dipolarophiles (R1 and                  distortion/interaction-activation strain analysis (Figure 11) to
R17) have asynchronous transition states (bond formation                       explore the origins of regioselectivity of the 1,3-dipolar
mainly at one center) and lower distortion energies at a given                 cycloadditions for the five asymmetrical alkenes in discussion.
forming bond length (Figure 9b). We interpret this as a result of              For R2, both the distortion and interaction energies determine
the stronger HOMO-LUMO interactions in the asynchronous                        the regioselectivity. TS2 is more asynchronous compared to
transition states: strong asynchronicity is related to lower                   TS2*. Therefore, it has lowered distortion energies for both
distortion energies and to much greater overlap at the shorter                 azide and alkene. Additionally, there are more favorable
forming bond.                                                                  interactions for TS2 due to the orbital interactions (the
   The interaction energies (Figure 9c) are higher for very                    combination between the terminal with the larger HOMO
electron-rich enamine (R1) and very electron-deficient dimethyl                 coefficient of alkene and the terminal with the larger LUMO+2
acetylenedicarboxylate (R21). Interaction energies involve                     coefficient of azide leads to more favorable orbital interaction),
more than the commonly cited frontier orbital interactions                     corresponding to a larger charge transfer (Table S2, entry 1).
represented in Figure 8, and we performed an energy                            Thus, the cycloaddition occurs via TS2 with higher selectivity.
decomposition analysis to determine the origins of the different                   For R8, R9, and R15, it is the interaction energy that
interaction energies (Figure S7). The interaction energy in                    determines the regioselectivity. There are stronger interactions
transition states (−8 to −14 kcal/mol) is a combination of Pauli               between phenyl azide and alkenes in the favorable transition
repulsion (70 to 100 kcal/mol), orbital interactions (around                   states (TS8, TS9, and TS15) compared to unfavorable ones
−40 kcal/mol), and electrostatic interactions (around −40 kcal/                (TS8*, TS9*, and TS15*) due to orbital interactions, and
mol). No clear pattern in the energy components was found.                     correspondingly, more significant charge transfer occurs in
The origin of these very large individual interactions is that                 favorable transition states (Table S2, entries 2−4). As for R17, it
covalent bonds are being formed and very large changes in                      is the distortion of azide that determines the regioselectivity.
                                                                        5799                                            https://doi.org/10.1021/acs.joc.1c00239
                                                                                                                             J. Org. Chem. 2021, 86, 5792−5804
The Journal of Organic Chemistry                                                  pubs.acs.org/joc                                               Article

Figure 10. Orbital interaction diagrams for the cycloadditions between phenyl azide and selected dipolarophiles. We added orbital coefficients for the
interacting orbitals. For phenyl azide, N1 is the nitrogen attached to the phenyl group, and N2 is the terminal-unsubstituted nitrogen. For
dipolarophiles, C3 is the carbon attached to the substituent, and C4 is the terminal-unsubstituted carbon. In order to justify that HOMO-3 and LUMO
+2 of phenyl azide are the effective orbitals involved in the orbital interactions between phenyl azide and dipolarophiles, we show the HOMO, HOMO-
1, and HOMO-2 as well as the LUMO and LUMO+1 of phenyl azide in the upper left corner of Figure 10 for comparison.

TS17 has a smaller distortion energy of azide compared to                     lectivity than the distortion energy. Therefore, the cycloaddition
TS17*. In terms of interaction energy, TS17* is more favorable                between phenyl azide and R17 occurs via TS17 selectively.
                                                                                 Based on the above analysis, we found that, in general, a
than TS17 due to orbital interactions (Table S2, entry 5).                    combination of interaction energies and distortion energies
However, the interaction energy contributes less to regiose-                  parallels those interactions predicated by the FMO model, and
                                                                       5800                                           https://doi.org/10.1021/acs.joc.1c00239
                                                                                                                           J. Org. Chem. 2021, 86, 5792−5804
The Journal of Organic Chemistry                                                      pubs.acs.org/joc                                                   Article

Figure 11. Distortion/interaction-activation strain (DIAS) analysis of the cycloaddition transition states to reveal the origins of regioselectivity. Free
energies were calculated at the ωB97X-D/aug-cc-pVTZ-CPCM(solvent)//ωB97X-D/6-31+G(d,p)-CPCM(solvent) level of theory. Fragment
distortion and interaction energies were calculated at the ωB97X-D level of theory with the aug-cc-pVTZ basis set, without the inclusion of solvation
energy corrections (black, activation energies; blue, distortion energies of azide; green, distortion energies of dipolarophiles; red, interaction energies).
Energies are in kcal/mol. The starred transition state is the regioisomeric transition state, which is unfavorable compared to the one without an asterisk.

the FMO models developed some time ago to rationalize and                         for the dipolarophiles with either low or high IPs, the HOMO-
predict 1,3-dipolar cycloaddition regioselectivity are supported                  LUMO gaps are small and distortion energies are small, at the
by this analysis.9                                                                same time that they are reactive nucleophiles or electrophiles.

■
                                                                                     Fukui’s frontier molecular orbital theory based only on
     CONCLUSIONS                                                                  HOMO-LUMO interactions is very successful, despite the
                                                                                  variety of interactions that influence reactivity.39 The FMO gap
As Sustmann showed in the 70s, the IP of dipolarophiles is a
                                                                                  is not only related to distortion energies but to electrostatic
powerful indicator of reactivity. A low IP (high HOMO)
                                                                                  interactions as well. As orbital interactions increase, charge
indicates a very nucleophilic molecule, an electron-rich
dipolarophile in our case, which reacts very rapidly with azide.                  transfer increases, causing electrostatic interactions between the
A high IP suggests lowering of the HOMO (and LUMO at the                          reactants to increase. Thus, the degree of electrostatic
same time); as the dipolarophile becomes more electrophilic, its                  interactions is also related to charge transfer, which is the result
reactivity with azide rises. We have also analyzed how the IP                     primarily of FMO interactions.
(minus of HOMO energy) relates to distortion, the FMO gap,                           Closed-shell repulsion is also related to FMO interactions.
closed-shell repulsions, and electrostatic and charge-transfer                    Strong HOMO-LUMO interactions between molecules are
interactions. All of these are involved in controlling reactivity.                generally accompanied by low closed-shell repulsions between
   Distortion energies are related to the HOMO-LUMO gap of                        HOMOs of the two molecules. A reduction in closed-shell
individual molecules through the “second-order Jahn−Teller                        repulsion increases reactivity.40 The primary example of this
effect”, which states that distortions are made easier by the                      relationship is found in the orbital symmetry relationships. An
mixing of the HOMO and LUMO of a molecule upon                                    orbital symmetry-allowed reaction has strong HOMO-LUMO
distortion; this becomes more effective when the HOMO-                             interactions between reactants because the HOMO of one
LUMO gap is small.38 A donor raises the HOMO energy (low                          molecule has the same symmetry as the LUMO of the other. For
IP), generally more than it raises the LUMO, which leads to a                     such a case, the HOMOs of the two molecules have opposite
small HOMO-LUMO gap. Alternatively, an electron acceptor                          symmetries and do not interact, and there is no destabilizing
lowers the HOMO energy (high IP), generally less than it lowers                   closed-shell repulsion between them. Donor and acceptor
the LUMO, which also narrows the HOMO-LUMO gap. Thus,                             substituents polarize the HOMO and LUMO of molecules in
                                                                           5801                                               https://doi.org/10.1021/acs.joc.1c00239
                                                                                                                                   J. Org. Chem. 2021, 86, 5792−5804
The Journal of Organic Chemistry

                                                                              ■
                                                                                  pubs.acs.org/joc                                                 Article

opposite directions, so there is always an opposite relationship                   ACKNOWLEDGMENTS
between stabilizing HOMO-LUMO and destabilizing HOMO-
                                                                              We are grateful to the Natural Science Foundation of Jiangsu
HOMO interactions.
                                                                              Province, China (BK20190505 to FL) and the National Science
   This intimate relationship between the IPs of a series of
                                                                              Foundation (CHE-1764328 to KNH) for financial support of
molecules and the LUMO energies, HOMO-LUMO gaps in a
                                                                              this research. D.S. is grateful to the Austrian Science Funds
molecule, and distortion, charge transfer, electrostatic, and
                                                                              (FWF, grant number J4216-N28) and the city of Vienna (H-
closed-shell repulsion effects accounts for the ability of IP to be
                                                                              331849/2018) for financial support. Calculations were
such a simple and valuable indicator of reactivity. As the power
                                                                              performed on the IDRE Hoffman2 cluster at the University of
of analysis increases, we have learned the profound prescience of
                                                                              California, Los Angeles, and the Vienna Scientific Cluster.

                                                                              ■
the Sustmann parabola.
   Looking back from the 2021 perspective at Sustmann’s
conclusion almost 50 years ago, we are struck by how insightful                    REFERENCES
and influential they were for the organic community, despite the                 (1) (a) Sustmann, R.; Trill, H. Substituent Effects in 1,3-Dipolar
lack of computers of much consequence when he published the                   Cycloadditions of Phenyl Azide. Angew. Chem. Int. Ed. Engl. 1972, 11,
paper (at least a billion times less powerful than today’s                    838−840. (b) Geittner, J.; Huisgen, R.; Sustmann, R. Kinetics of 1,3-
computers). The real chemical reaction system is complex, and                 Dipolar Cycloaddition Reactions of Diazomethane; A Correlation with
one parameter can never be quantitatively related to reactivity.              HOMO-LUMO Energies. Tetrahedron Letters. 1977, 18, 881−884.
Nevertheless, chemists appreciate models that give a qualitative              (c) Sustmann, R. A Simple Model for Substituent Effects in
                                                                              Cycloaddition Reactions. I. 1,3-Dipolar Cycloadditions. Tetrahedron
guide to reactivity. Nowadays, armed with sophisticated
                                                                              Letters. 1971, 12, 2717−2720.
methods of theoretical calculation and analysis, we can provide                 (2) Dewar, M. J. S. The Molecular Orbital Theory of Organic Chemistry,
intimate and complicated details about mechanisms and                         McGraw-Hill, 1969.
reactivity. Nevertheless, we acknowledge and admire Sus-                        (3) For some early pioneering experimental studies, see: (a) Smith, L.
tmann’s insight that the IP of dipolarophiles is a good guide to              I. Aliphatic Diazo Compounds, Nitrones, and Structurally Analogous
reactivities of dipolarophiles toward ambiphilic azides.                      Compounds. Systems Capable of Undergoing 1,3-Additions. Chem.

■
                                                                              Rev. 1938, 23, 193−285. (b) Huisgen, R. Kinetics and Mechanism of
     ASSOCIATED CONTENT                                                       1,3-Dipolr Cycloadditions. Angew. Chem. Int. Ed. Engl. 1963, 2, 633−
                                                                              645.
*
sı Supporting Information                                                       (4) (a) Bischof, P.; Hashmall, J. A.; Heilbronner, E.; Hornung, V.
The Supporting Information is available free of charge at                     Photoelektronspektroskopische Bestimmunǵ der Wechselwirkunǵ
https://pubs.acs.org/doi/10.1021/acs.joc.1c00239.                             zwischen nicht-konjuǵierten Doppelbindunǵen [1] Vorläufige Mittei-
                                                                              lung. Helv. Chim. Acta. 1969, 52, 1745−1749. (b) Bischof, P.;
        Optimized geometries and energies of all computed                     Heilbronner, E. Photoelektron-Spektren von Cycloalkenen und
        species (PDF)                                                         Cycloalkadienen. Vorläufige Mitteilung. Helv. Chim. Acta. 1970, 53,

■
                                                                              1677−1682.
                                                                                (5) (a) Sustmann, R.; Schubert, R. Photoelektronenspektroskopische
     AUTHOR INFORMATION                                                       Bestimmung von Substituenten-Effekten I, Substituierte Butadiene.
                                                                              Tetrahedron Letters. 1972, 13, 2739−2742. (b) Sustmann, R.; Trill, H.
Corresponding Authors                                                         Photoelektronenspektroskopische Bestimmung von Substituenten-
    Fang Liu − College of Sciences, Nanjing Agricultural University,          Effekten II. α,β-ungesättigte Carbonester. Tetrahedron Letters. 1972,
      Nanjing 210095, China; Email: acialiu@njau.edu.cn                       13, 4271−4274.
    Kendall N. Houk − Department of Chemistry and Biochemistry,                 (6) (a) Domelsmith, N.; Houk, K. N. Photoelectron Spectra of
      University of California, Los Angeles, California 90095-1569,           Cyclopentanone and Cyclohexanone Enamines. Tetrahedron Letters.
      United States; orcid.org/0000-0002-8387-5261;                           1977, 18, 1981−1984. (b) McAlduff, E. J.; Caramella, P.; Houk, K. N.
      Email: houk@chem.ucla.edu                                               Photoelectron Spectra of 3-Substituted Cyclopentenes. Correlations
                                                                              between Ionization Potentials and Cycloaddition Regioselectivity. J.
Authors                                                                       Am. Chem. Soc. 1978, 100, 105−110.
   Pan-Pan Chen − Department of Chemistry and Biochemistry,                     (7) (a) Masclet, P.; Grosjean, D.; Mouvier, G.; Dubois, J. Alkene
                                                                              Ionization Potentials: Part I: Quantitative Determination of Alkyl
     University of California, Los Angeles, California 90095-1569,
                                                                              Group Structural Effects. Journal of Electron Spectroscopy and Related
     United States                                                            Phenomena. 1973, 2, 225−237. (b) Rabalais, J. W.; Colton, R. J.
   Pengchen Ma − Department of Chemistry and Biochemistry,                    Electronic Interaction between the Phenyl Group and Its Unsaturated
     University of California, Los Angeles, California 90095-1569,            Substituents. Journal of Electron Spectroscopy & Related Phenomena
     United States                                                            1972, 1, 83−99.
   Xue He − College of Sciences, Nanjing Agricultural University,               (8) Baker, A. D.; Betteridge, D. Photoelectron Spectroscopy. Chemical
     Nanjing 210095, China                                                    and Analytical Aspects; Pergamon Press: Oxford, 1972.
   Dennis Svatunek − Department of Chemistry and Biochemistry,                  (9) (a) Houk, K. N. Regioselectivity and Reactivity in the 1,3-Dipolar
     University of California, Los Angeles, California 90095-1569,            Cycloadditions of Diazonium Betaines (Diazoalkanes, Azides, and
     United States; orcid.org/0000-0003-1101-2376                             Nitrous Oxide). J. Am. Chem. Soc. 1972, 94, 8953−8955. (b) Houk, K.
                                                                              N.; Sims, J.; Duke, R. E.; Strozier, R. W.; George, J. K. Frontier
Complete contact information is available at:                                 Molecular Orbitals of 1,3 Dipoles and Dipolarophiles. J. Am. Chem. Soc.
https://pubs.acs.org/10.1021/acs.joc.1c00239                                  1973, 95, 7287−7301. (c) Houk, K. N.; Sims, J.; Watts, C. R.; Luskus,
                                                                              L. J. Origin of Reactivity, Regioselectivity, and Periselectivity in 1,3-
Author Contributions                                                          Dipolar Cycloadditions. J. Am. Chem. Soc. 1973, 95, 7301−7315.
§                                                                               (10) (a) Ess, D. H.; Houk, K. N. Distortion/Interaction Energy
 P.-P.C. and P.M. contributed equally to this manuscript.
                                                                              Control of 1,3-Dipolar Cycloaddition Reactivity. J. Am. Chem. Soc.
Notes                                                                         2007, 129, 10646−10647. (b) Ess, D. H.; Jones, G. O.; Houk, K. N.
The authors declare no competing financial interest.                           Transition States of Strain-Promoted Metal-Free Click Chemistry: 1,3-

                                                                       5802                                             https://doi.org/10.1021/acs.joc.1c00239
                                                                                                                             J. Org. Chem. 2021, 86, 5792−5804
The Journal of Organic Chemistry                                                        pubs.acs.org/joc                                                   Article

Dipolar Cycloadditions of Phenyl Azide and Cyclooctynes. Org. Lett.                 Heyd, J. J.; Brothers, E. N.; Kudin, K. N.; Staroverov, V. N.; Keith, T. A.;
2008, 10, 1633−1636. (c) Schoenebeck, F.; Ess, D. H.; Jones, G. O.;                 Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A. P.; Burant, J.
Houk, K. N. Reactivity and Regioselectivity in 1,3-Dipolar Cyclo-                   C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Millam, J. M.; Klene, M.;
additions of Azides to Strained Alkynes and Alkenes: A Computational                Adamo, C.; Cammi, R.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.;
Study. J. Am. Chem. Soc. 2009, 131, 8121−8133. (d) Gordon, C. G.;                   Farkas, O.; Foresman, J. B.; Fox, D. J. Gaussian 16, Rev. A.03,
Mackey, J. L.; Jewett, J. C.; Sletten, E. M.; Houk, K. N.; Bertozzi, C. R.          Wallingford, CT, 2016.
Reactivity of Biarylazacyclooctynones in Copper-Free Click Chemistry.                 (17) Chai, J.-D.; Head-Gordon, M. Long-Range Corrected Hybrid
J. Am. Chem. Soc. 2012, 134, 9199−9208. (e) Bickelhaupt, F. M.; Houk,               Density Functionals with Damped Atom−Atom Dispersion Correc-
K. N. Analyzing Reaction Rates with the Distortion/Interaction-                     tions. Phys. Chem. Chem. Phys. 2008, 10, 6615−6620.
Activation Strain Model. Angew. Chem. Int. Ed. 2017, 56, 10070−                       (18) (a) Grimme, S. Semiempirical Hybrid Density Functional with
10086. (f) Hamlin, T. A.; Svatunek, D.; Yu, S.; Ridder, L.; Infante, I.;            Perturbative Second-order Correlation. J. Chem. Phys. 2006, 124,
Visscher, L.; Bickelhaupt, F. M. Elucidating the Trends in Reactivity of            No. 034108. (b) Grimme, S. Semiempirical GGA-Type Density
Aza-1,3-Dipolar Cycloadditions. Eur. J. Org. Chem. 2019, 378−386.                   Functional Constructed with A Long-Range Dispersion Correction. J.
  (11) (a) Domingo, L. R. Molecular Electron Density Theory: A                      Comput. Chem. 2006, 27, 1787−1799.
Modern View of Reactivity in Organic Chemistry. Molecules 2016, 21,                   (19) (a) Dunning, T. H. Gaussian-Basis Sets for Use in Correlated
1319. Applications to azide cycloadditions: (b) Ben El Ayouchia, H.;                Molecular Calculations. I. The Atoms Boron through Neon and
Lahoucine, B.; Anane, H.; Ríos-Gutiérrez, M.; Domingo, L. R.; Stiriba,              Hydrogen. J. Chem. Phys. 1989, 90, 1007−1023. (b) Kendall, R. A.;
S.-E. Experimental and Theoretical MEDT Study of the Thermal [3+2]                  Dunning Jr, T. H.; Harrison, R. J. Electron Affinities of the First-Row
Cycloaddition Reactions of Aryl Azides with Alkyne Derivatives.                     Atoms Revisited. Systematic Basis Sets and Wave Functions. J. Chem.
ChemistrySelect 2018, 3, 1215−1223. Domingo, L. R.; Acharjee, N.                    Phys. 1992, 96, 6796−6806.
Unravelling the Strain-Promoted [3+2] Cycloadditi on Reactions of                     (20) Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. Energies,
Phenyl Azide with Cycloalkynes from the Molecular Electron Density                  Structures, and Electronic Properties of Molecules in Solution with
Theory Perspective. New J. Chem. 2020, 44, 13633-13643.                             the C-PCM Solvation Model. J. Comput. Chem. 2003, 24, 669−681.
  (12) Freindorf, M.; Sexton, T.; Kraka, E.; Cremer, D. The Mechanism                 (21) (a) Lu, T.; Chen, F. Multiwfn: A Multifunctional Wavefunction
of the Cycloaddition Reaction of 1,3-Dipole Molecules with Acetylene:               Analyzer. J. Comput. Chem. 2012, 33, 580−592. (b) Lu, T.; Chen, F.
An Investigation with the Unified Reaction Valley Approach. Theoret.                Quantitative Analysis of Molecular Surface Based on Improved
Chem. Acc. 2014, 133, 1423.                                                         Marching Tetrahedra Algorithm. J. Mol. Graphics Model. 2012, 38,
  (13) (a) Riplinger, C.; Neese, F. An Efficient and Near Linear Scaling            314−323.
Pair Natural Orbital Based Local Coupled Cluster Method. J. Chem.                     (22) Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual Molecular
Phys. 2013, 138, No. 034106. (b) Riplinger, C.; Sandhoefer, B.; Hansen,             Dynamics. J. Mol. Graphics. 1996, 14, 33−38. VMD Official website.
A.; Neese, F. Natural Triple Excitations in Local Coupled Cluster                   http://www.ks.uiuc.edu/Research/ vmd/.
Calculations with Pair Natural Orbitals. J. Chem. Phys. 2013, 139,                    (23) Svatunek, D.; Houk, K. N. autoDIAS: A Python Tool for an
134101. (c) Liakos, D. G.; Neese, F. Is It Possible To Obtain Coupled               Automated Distortion/Interaction Activation Strain Analysis. J.
Cluster Quality Energies at near Density Functional Theory Cost?                    Comput. Chem. 2019, 40, 2509−2515.
Domain-Based Local Pair Natural Orbital Coupled Cluster vs. Modern                    (24) te Velde, G.; Bickelhaupt, F. M.; Baerends, E. J.; Fonseca Guerra,
Density Functional Theory. J. Chem. Theory Comput. 2015, 11, 4054−                  C.; van Gisbergen, S. J. A.; Snijders, J. G.; Ziegler, T. Chemistry with
4063.                                                                               ADF. J. Comput. Chem. 2001, 22, 931−967.
  (14) (a) Breugst, M.; Huisgen, R.; Reissig, H.-U. Regioselective 1,3-               (25) Baerends, E. J.; Ellis, D. E.; Ros, P. Self-Consistent Molecular
Dipolar Cycloadditions of Diazoalkanes with Heteroatom-Substituted                  Hartree-Fock-Slater Calculations I. The Computational Procedure.
Alkynes: Theory and Experiment. Eur. J. Org. Chem. 2018, 2477−2485.                 Chem. Phys. 1973, 2, 41−51.
(b) Breugst, M.; Reissig, H.-U. The Huisgen Reaction: Milestones of                   (26) Sun, X.; Soini, T. M.; Poater, J.; Hamlin, T. A.; Bickelhaupt, F. M.
the 1,3-Dipolar Cycloaddition. Angew. Chem. Int. Ed. 2020, 59, 12293−               PyFrag 2019-Automating the Exploration and Analysis of Reaction
12307.                                                                              Mechanisms. J. Comput. Chem. 2019, 40, 2227−2233.
  (15) For some selected computational studies on the reactivity of                   (27) (a) Foster, J. P.; Weinhold, F. Natural Hybrid Orbitals. J. Am.
azides in 1,3-dipolar cycloadditions, see: (a) Hamlin, T. A.;                       Chem. Soc. 1980, 102, 7211−7218. (b) Reed, A. E.; Weinstock, R. B.;
Levandowski, B. J.; Narsaria, A. K.; Houk, K. N.; Bickelhaupt, F. M.                Weinhold, F. Natural Population Analysis. J. Chem. Phys. 1985, 83,
Structural Distortion of Cycloalkynes Influences Cycloaddition Rates                735−746.
both by Strain and Interaction Energies. Chem. Eur. J. 2019, 25, 6342−                (28) Legault, C. Y. CYLview; version 1.0b; Université de Sherbrooke:
6348. (b) Dommerholt, J.; van Rooijen, O.; Borrmann, A.; Fonseca                    2009 (http://www.cylview.org).
Guerra, C.; Bickelhaupt, F. M.; van Delft, F. L. Highly Accelerated                   (29) We calculated the rate constants from the activation barriers
Inverse Electron-Demand Cycloaddition of Electron-Deficient Azides                  based on Eyring Equation, which follows from the transition state
with Aliphatic Cyclooctynes. Nat. Commun. 2014, 5, 5378. (c) Lopez, S.              theory, also known as activated-complex theory.
A.; Munk, M. E.; Houk, K. N. Mechanisms and Transition States of 1,3-                 (30) Huisgen, R.; Szeimie, G.; Möbius, L. 1.3-Dipolare Cyclo-
Dipolar Cycloadditions of Phenyl Azide with Enamines: A Computa-                    additionen, XXXII. Kinetik der Additionen organischer Azide an CC-
tional Analysis. J. Org. Chem. 2013, 78, 1576−1582. (d) Lopez, S. A.;               Mehrfachbindungen. Chem. Ber. 1967, 100, 2494−2507.
Houk, K. N. Alkene Distortion Energies and Torsional Effects Control                  (31) Koopmans, T. Ü ber die Zuordnung von Wellenfunktionen und
Reactivities, and Stereoselectivities of Azide Cycloadditions to                    Eigenwerten zu den einzelnen Elektronen eines Atoms. Physica. 1934,
Norbornene and Substituted Norbornenes. J. Org. Chem. 2013, 78,                     1, 104−113.
1778−1783.                                                                            (32) The exact dipolarophile used in the experiment is 1-
  (16) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;              ethoxycyclopentene.30 In the calculation, 1-methoxycyclopentene is
Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson, G.              substituted for 1-ethoxycyclopentene in order to simplify the
A.; Nakatsuji, H.; Li, X.; Caricato, M.; Marenich, A. V.; Bloino, J.;               calculation model.
Janesko, B. G.; Gomperts, R.; Mennucci, B.; Hratchian, H. P.; Ortiz, J.               (33) Liu, F.; Liang, Y.; Houk, K. N. Theoretical Elucidation of the
V.; Izmaylov, A. F.; Sonnenberg, J. L.; Williams; Ding, F.; Lipparini, F.;          Origins of Substituent and Strain Effects on the Rates of Diels−Alder
Egidi, F.; Goings, J.; Peng, B.; Petrone, A.; Henderson, T.; Ranasinghe,            Reactions of 1,2,4,5-Tetrazines. J. Am. Chem. Soc. 2014, 136, 11483−
D.; Zakrzewski, V. G.; Gao, J.; Rega, N.; Zheng, G.; Liang, W.; Hada,               11493.
M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.;                      (34) (a) Agard, N. J.; Prescher, J. A.; Bertozzi, C. R. A Strain-
Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Throssell,               Promoted [3 + 2] Azide−Alkyne Cycloaddition for Covalent
K.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M. J.;           Modification of Biomolecules in Living Systems. J. Am. Chem. Soc.

                                                                             5803                                               https://doi.org/10.1021/acs.joc.1c00239
                                                                                                                                     J. Org. Chem. 2021, 86, 5792−5804
The Journal of Organic Chemistry                                                     pubs.acs.org/joc                              Article

2004, 126, 15046−15047. (b) Lutz, J.-F. Copper-Free Azide−Alkyne
Cycloadditions: New Insights and Perspectives. Angew. Chem., Int. Ed.
2008, 47, 2182−2184. (c) Ning, X.; Guo, J.; Wolfert, M. A.; Boons, G.-
J. Visualizing Metabolically Labeled Glycoconjugates of Living Cells by
Copper-Free and Fast Huisgen Cycloadditions. Angew. Chem., Int. Ed.
2008, 47, 2253−2255. (d) Patterson, D. M.; Nazarova, L. A.; Xie, B.;
Kamber, D. N.; Prescher, J. A. Functionalized Cyclopropenes as
Bioorthogonal Chemical Reporters. J. Am. Chem. Soc. 2012, 134,
18638−18643. (e) Yang, J.; Š ečkutė, J.; Cole, C. M.; Devaraj, N. K.
Live-Cell Imaging of Cyclopropene Tags with Fluorogenic Tetrazine
Cycloadditions. Angew. Chem., Int. Ed. 2012, 51, 7476−7479. (f) Liu,
F.; Liang, Y.; Houk, K. N. Bioorthogonal Cycloadditions: Computa-
tional Analysis with the Distortion/Interaction Model and Prediction of
Reactivities. Acc. Chem. Res. 2017, 50, 2297−2308.
  (35) Wang, J.; Yang, B.; Cool, T. A.; Hansen, N.; Kasper, T. Near-
Threshold Absolute Photoionization Cross-Sections of Some Reaction
Intermediates in Combustion. Int. J. Mass Spectrom. 2008, 269, 210−
220.
  (36) Liang, Y.; Mackey, J. L.; Lopez, S. A.; Liu, F.; Houk, K. N. Control
and Design of Mutual Orthogonality in Bioorthogonal Cycloadditions.
J. Am. Chem. Soc. 2012, 134, 17904−17907.
  (37) Himo, F.; Lovell, T.; Hilgraf, R.; Rostovtsev, V. V.; Noodleman,
L.; Sharpless, K. B.; Fokin, V. V. Copper(I)-Catalyzed Synthesis of
Azoles. DFT Study Predicts Unprecedented Reactivity and Inter-
mediates. J. Am. Chem. Soc. 2005, 127, 210−216.
  (38) (a) Bader, R. F. W. An Interpretation of Potential Interaction
Constants in Terms of Low-Lying Excited States. Molecular Physics.
1960, 3, 137−151. (b) Pearson, R. G. Concerning Jahn-Teller Effects.
Proc. Nat. Acad. Sci. USA. 1975, 72, 2104−2106.
  (39) (a) Fukui, K.; Yonezawa, T.; Shingu, H. A Molecular Orbital
Theory of Reactivity in Aromatic Hydrocarbons. J. Chem. Phys. 1952,
20, 722−725. (b) Fukui, K. Recognition of Stereochemical Paths by
Orbital Interaction. Acc. Chem. Res. 1971, 4, 57−64. (c) Fleming, I.
Frontier Orbitals and Organic Chemical Reactions. London: Wiley.
1978, 24−109. ISBN 0-471-01819-8.
  (40) (a) Hamlin, T. A.; Fernández, I.; Bickelhaupt, F. M. How
Dihalogens Catalyze Michael Addition Reactions. Angew. Chem. Int. Ed.
2019, 58, 8922−8926. (b) Vermeeren, P.; Brinkhuis, F.; Hamlin, T. A.;
Bickelhaupt, F. M. How Alkali Cations Catalyze Aromatic Diels-Alder
Reactions. Chem Asian J. 2020, 15, 1167−1174. (c) Vermeeren, P.;
Hamlin, T. A.; Fernández, I.; Bickelhaupt, F. M. How Lewis Acids
Catalyze Diels−Alder Reactions. Angew. Chem. Int. Ed. 2020, 59, 6201−
6206.

■    NOTE ADDED AFTER ISSUE PUBLICATION
This article was initially published with an incorrect copyright
statement and was corrected on or around May 5, 2021.

                                                                              5804                      https://doi.org/10.1021/acs.joc.1c00239
                                                                                                             J. Org. Chem. 2021, 86, 5792−5804
You can also read